Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 30

CHAPTER 1

INTRODUCTION

An organic solar cell or plastic solar cell is a type of photovoltaic that uses
organic electronics, a branch of electronics that deals with conductive organic
polymers or small organic molecules, for light absorption and charge transport
to produce electricity from sunlight by the photovoltaic effect. Most organic
photovoltaic cells are polymer solar cells.

The molecules used in organic solar cells are solution-processable at high


throughput and are cheap, resulting in low production costs to fabricate a large
volume. Combined with the flexibility of organic molecules, organic solar cells
are potentially cost-effective for photovoltaic applications. Molecular
engineering (e.g., changing the length and functional group of polymers) can
change the band gap, allowing for electronic tunability. The optical absorption
coefficient of organic molecules is high, so a large amount of light can be
absorbed with a small amount of materials, usually on the order of hundreds of
nanometers. The main disadvantages associated with organic photovoltaic cells
are low efficiency, low stability and low strength compared to inorganic
photovoltaic cells such as silicon solar cells.

Compared to silicon-based devices, polymer solar cells are lightweight


(which is important for small autonomous sensors), potentially disposable and
inexpensive to fabricate (sometimes using printed electronics), flexible,
customizable on the molecular level and potentially have less adverse
environmental impact. Polymer solar cells also have the potential to exhibit
transparency, suggesting applications in windows, walls, flexible electronics,

1
etc. An example device is shown in Fig. 1. The disadvantages of polymer solar
cells are also serious: they offer about 1/3 of the efficiency of hard materials,
and experience substantial photochemical degradation.

Polymer solar cells inefficiency and stability problems, combined with their
promise of low costs and increased efficiency made them a popular field in
solar cell research. As of 2015, polymer solar cells were able to achieve over
10% efficiency via a tandem structure. In 2018, a record breaking efficiency for
organic photovoltaics of 17.3% was reached via tandem structure.

Figure 1.1

2
CHAPTER 2
WORKING PRINCIPLE
A photovoltaic cell is a specialized semiconductor diode that converts light
into direct current (DC) electricity. Depending on the band gap of the light-
absorbing material, photovoltaic cells can also convert low-energy, infrared (IR)
or high-energy, ultraviolet (UV) photons into DC electricity. A common
characteristic of both the small molecules and polymers used as the light-
absorbing material in photovoltaics is that they all have large conjugated
systems. A conjugated system is formed where carbon atoms covalently bond
with alternating single and double bonds. These hydrocarbons' electrons pz
orbitals delocalize and form a delocalized bonding π orbital with a π*
antibonding orbital. The delocalized π orbital is the highest occupied molecular
orbital, and the π* orbital is the lowest unoccupied molecular orbital. In organic
semiconductor physics, It takes the role of the valence band while the LUMO
serves as the conduction band. The energy separation between the HOMO and
LUMO energy levels is considered the band gap of organic electronic materials
and is typically in the range of 1–4 eV.

All light with energy greater than the band gap of the material can be
absorbed, though there is a trade-off to reducing the band gap as photons
absorbed with energies higher than the band gap will thermally give off their
excess energy, resulting in lower voltages and power conversion efficiencies.
When these materials absorb a photon, an excited state is created and confined
to a molecule or a region of a polymer chain. The excited state can be regarded
as an exciton, or an electron-hole pair bound together by electrostatic
interactions. In photovoltaic cells, excitons are broken up into free electron-hole

3
pairs by effective fields. The effective fields are set up by creating a
heterojunction between two dissimilar materials. In organic photovoltaics,
effective fields break up excitons by causing the electron to fall from the
conduction band of the absorber to the conduction band of the acceptor
molecule. It is necessary that the acceptor material has a conduction band edge
that is lower than that of the absorber material.

Polymer solar cells usually consist of an electron- or hole-blocking layer on


top of an indium tin oxide (ITO) conductive glass followed by electron donor
and an electron acceptor (in the case of bulk heterojunction solar cells), a hole
or electron blocking layer, and metal electrode on top. The nature and order of
the blocking layers – as well as the nature of the metal electrode – depends on
whether the cell follows a regular or an inverted device architecture. In an
inverted cell, the electric charges exit the device in the opposite direction as in a
normal device because the positive and negative electrodes are reversed.
Inverted cells can utilize cathodes out of a more suitable material; inverted
OPVs enjoy longer lifetimes than regularly structured OPVs, and they usually
show higher efficiencies compared with the conventional counterparts.

In bulk heterojunction polymer solar cells, light generates excitons.


Subsequent charge separation in the interface between an electron donor and
acceptor blend within the device's active layer. These charges then transport to
the device's electrodes where the charges flow outside the cell, perform work
and then re-enter the device on the opposite side. The cell's efficiency is limited
by several factors, especially non-geminate recombination. Hole mobility leads
to faster conduction across the active layer.

4
Organic photovoltaics are made of electron donor and electron acceptor
materials rather than semiconductor p-n junctions. The molecules forming the
electron donor region of organic PV cells, where exciton electron-hole pairs are
generated, are generally conjugated polymers possessing delocalized π electrons
that result from carbon p orbital hybridization. These π electrons can be excited
by light in or near the visible part of the spectrum from the molecule's highest
occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital
(LUMO), denoted by a π -π* transition. The energy bandgap between these
orbitals determines which wavelength(s) of light can be absorbed.

Unlike in an inorganic crystalline PV cell material, with its band structure


and delocalized electrons, excitons in organic photovoltaics are strongly bound
with an energy between 0.1 and 1.4 eV. This strong binding occurs because
electronic wave functions in organic molecules are more localized, and
electrostatic attraction can thus keep the electron and hole together as an
exciton. The electron and hole can be dissociated by providing an interface
across which the chemical potential of electrons decreases. The material that
absorbs the photon is the donor, and the material acquiring the electron is called
the acceptor. The polymer chain is the donor and the fullerene is the acceptor.
Even after dissociation, the electron and hole may still be joined as a "geminate
pair", and an electric field is then required to separate them. The electron and
hole must be collected at contacts. If charge carrier mobility is insufficient, the
carriers will not reach the contacts, and instead recombine at trap sites or remain
in the device as undesirable space charges that oppose the flow of new carriers.
The latter problem can occur if electron and hole mobilities are not matched. In
that case, space-charge limited photocurrent (SCLP) hampers device
performance.

5
Organic photovoltaics can be fabricated with an active polymer and a
fullerene-based electron acceptor. Illumination of this system by visible light
leads to electron transfer from the polymer to a fullerene molecule. As a result,
the formation of a photoinduced quasiparticle, or polaron (P+), occurs on the
polymer chain and the fullerene becomes a radical anion. Polarons are highly
mobile and can diffuse away.

6
CHAPTER 3
JUNCTION TYPES
The simplest organic PV device features a planar heterojunction. A film of
organic active material (polymer or small molecule), of electron donor or
electron acceptor type is sandwiched between contacts. Excitons created in the
active material may diffuse before recombining and separate, hole and electron
diffusing to its specific collecting electrode. Because charge carriers have
diffusion lengths of just 3–10 nm in typical amorphous organic semiconductors,
planar cells must be thin, but the thin cells absorb light less well. Bulk
heterojunctions (BHJs) address this shortcoming. In a BHJ, a blend of electron
donor and acceptor materials is cast as a mixture, which then phase-separates.
Regions of each material in the device are separated by only several
nanometers, a distance suited for carrier diffusion. BHJs require sensitive
control over materials morphology on the nanoscale. Important variables
include materials, solvents and the donor-acceptor weight ratio.

The next logical step beyond BHJs are ordered nanomaterials for solar cells,
or ordered heterojunctions (OHJs). OHJs minimize the variability associated
with BHJs. OHJs are generally hybrids of ordered inorganic materials and
organic active regions. For example, a photovoltaic polymer can be deposited
into pores in a ceramic such as TiO2. Since holes still must diffuse the length of
the pore through the polymer to a contact, OHJs suffer similar thickness
limitations. Mitigating the hole mobility bottleneck is key to further enhancing
device performance of OHJ's.

7
3.1 Single layer
Single layer organic photovoltaic cells are the simplest form. These cells are
made by sandwiching a layer of organic electronic materials between two
metallic conductors, typically a layer of indium tin oxide (ITO) with high work
function and a layer of low work function metal such as Aluminum, Magnesium
or Calcium. The basic structure of such a cell is illustrated in.

The difference of work function between the two conductors sets up an


electric field in the organic layer. When the organic layer absorbs light,
electrons will be excited to the LUMO and leave holes in the HOMO, thereby
forming excitons. The potential created by the different work functions helps to
split the exciton pairs, pulling electrons to the positive electrode (an electrical
conductor used to make contact with a non-metallic part of a circuit) and holes
to the negative electrode.

Figure 3.1

Conjugated polymers were also used in this type of photovoltaic cell. One
device used polyacetylene as the organic layer, with Al and graphite,
producing an open-circuit voltage of 0.3 V and a charge collection efficiency of
0.3%. An Al/poly(3-nethyl-thiophene)/Pt cell had an exteral quantum yield

8
of 0.17%, an open-circuit voltage of 0.4 V and a fill factor of 0.3. An
ITO/PPV/Al cell showed an open-circuit voltage of 1 V and a power conversion
efficiency of 0.1% under white-light illumination.

Single layer organic solar cells do not work well. They have low quantum
efficiencies (<1%) and low power conversion efficiencies (<0.1%). A major
problem with them is that the electric field resulting from the difference
between the two conductive electrodes is seldom sufficient to split the excitons.
Often the electrons recombine with the holes without reaching the electrode.

3.2 Bilayer
Bilayer cells contain two layers in between the conductive electrodes. The
two layers have different electron affinity and ionization energies, therefore
electrostatic forces are generated at the interface between the two layers. Light
must create excitons in this small charged region for an efficient charge
separation and collecting. The materials are chosen to make the differences
large enough that these local electric fields are strong, which splits excitons
much more efficiently than single layer photovoltaic cells. The layer with higher
electron affinity and ionization potential is the electron acceptor, and the other
layer is the electron donor. This structure is also called a planar donor-acceptor
heterojunction.

9
Figure 3.2

C60 has high electron affinity, making it a good acceptor. A C60/MEH-PPV


double layer cell had a relatively high fill factor of 0.48 and a power conversion
efficiency of 0.04% under monochromatic illumination. PPV/C60 cells
displayed a monochromatic external quantum efficiency of 9%, a power
conversion efficiency of 1% and a fill factor of 0.48.

Perylene derivatives display high electron affinity and chemical stability. A


layer of copper phthalocyanine (CuPc) as electron donor and perylene
tetracarboxylic derivative as electron acceptor, fabricating a cell with a fill
factor as high as 0.65 and a power conversion efficiency of 1% under simulated
AM2 illumination. Halls et al. fabricated a cell with a layer of
bis(phenethylimido) perylene over a layer of PPV as the electron donor. This
cell had peak external quantum efficiency of 6% and power conversion
efficiency of 1% under monochromatic illumination, and a fill factor of up to
0.6.

3.3 Bulk heterojunction


Bulk heterojunctions have an absorption layer consisting of a nanoscale
blend of donor and acceptor materials. The domain sizes of this blend are on the

10
order of nanometers, allowing for excitons with short lifetimes to reach an
interface and dissociate due to the large donor-acceptor interfacial area.
However, efficient bulk heterojunctions need to maintain large enough domain
sizes to form a percolating network that allows the donor materials to reach the
hole transporting electrode (Electrode 1 in Fig. 7) and the acceptor materials to
reach the electron transporting electrode (Electrode 2). Without this percolating
network, charges might be trapped in a donor or acceptor rich domain and
undergo recombination. Bulk heterojunctions have an advantage over layered
photoactive structures because they can be made thick enough for effective
photon absorption without the difficult processing involved in orienting a
layered structure while retaining similar level of performances.

Bulk heterojunctions are most commonly created by forming a solution


containing the two components, casting (e.g., drop casting and spin coating) and
then allowing the two phases to separate, usually with the assistance of an
annealing step. The two components will self-assemble into an interpenetrating
network connecting the two electrodes. They are normally composed of a
conjugated molecule based donor and fullerene based acceptor. The
nanostructural morphology of bulk heterojunctions tends to be difficult to
control, but is critical to photovoltaic performance.

After the capture of a photon, electrons move to the acceptor domains, then
are carried through the device and collected by one electrode, and holes move in
the opposite direction and collected at the other side. If the dispersion of the two
materials is too fine, it will result in poor charge transfer through the
layer.

11
Most bulk heterojunction cells use two components, although three-
component cells have been explored. The third component, a secondary p-type
donor polymer, acts to absorb light in a different region of the solar spectrum.
This in theory increases the amount of absorbed light. These ternary cells
operate through one of three distinct mechanisms: charge transfer, energy
transfer or parallel-linkage.

In charge transfer, both donors contribute directly to the generation of free


charge carriers. Holes pass through only one donor domain before collection at
the anode. In energy transfer, only one donor contributes to the production of
holes. The second donor acts solely to absorb light, transferring extra energy to
the first donor material. In parallel linkage, both donors produce excitons
independently, which then migrate to their respective donor/acceptor interfaces
and dissociate.

12
CHAPTER 4
PRODUCTION

If one material is more soluble in the solvent than the other, it will deposit
first on top of the substrate, causing a concentration gradient through the film.
This has been demonstrated for poly-3-hexyl thiophene (P3HT), phenyl-C61-
butyric acid methyl ester (PCBM) devices where the PCBM tends to
accumulate towards the device's bottom upon spin coating from ODCB
solutions This effect is seen because the more soluble component tends to
migrate towards the "solvent rich" phase during the coating procedure,
accumulating the more soluble component towards the film's bottom, where the
solvent remains longer. The thickness of the generated film affects the phases
segregation because the dynamics of crystallization and precipitation are
different for more concentrated solutions or faster evaporation rates (needed to
build thicker devices). Crystalline P3HT enrichment closer to the hole-
collecting electrode can only be achieved for relatively thin (100 nm)
P3HT/PCBM layers.
The gradients in the initial morphology are then mainly generated by the
solvent evaporation rate and the differences in solubility between the donor and
acceptor inside the blend. This dependence on solubility has been clearly
demonstrated using fullerene derivatives and P3HT. When using solvents
which evaporate at a slower rate (as chlorobenzene (CB) or dichlorobenzene
(DCB)) you can get larger degrees of vertical separation or aggregation while
solvents that evaporate quicker produce a much less effective vertical
separation. Larger solubility gradients should lead to more effective vertical

13
separation while smaller gradients should lead to more homogeneous films.
These two effects were verified on P3HT:PCBM solar cells.

The solvent evaporation speed as well as posterior solvent vapor or thermal


annealing procedures were also studied. Blends such as P3HT:PCBM seem
to benefit from thermal annealing procedures, while others, such as
PTB7:PCBM, seem to show no benefit. In P3HT the benefit seems to come
from an increase of crystallinity of the P3HT phase which is generated through
an expulsion of PCBM molecules from within these domains. This has been
demonstrated through studies of PCBM miscibility in P3HT as well as domain
composition changes as a function of annealing times.

The above hypothesis based on miscibility does not fully explain the
efficiency of the devices as solely pure amorphous phases of either donor or
acceptor materials never exist within bulk heterojunction devices. A 2010 paper
suggested that current models that assume pure phases and discrete interfaces
might fail given the absence of pure amorphous regions. Since current models
assume phase separation at interfaces without any consideration for phase
purity, the models might need to be changed.

The thermal annealing procedure varies depending on precisely when it is


applied. Since vertical species migration is partly determined by the surface
tension between the active layer and either air or another layer, annealing before
or after the deposition of additional layers (most often the metal cathode)
affects the result. In the case of P3HT:PCBM solar cells vertical migration is
improved when cells are annealed after the deposition of the metal cathode.

14
Donor or acceptor accumulation next to the adjacent layers might be
beneficial as these accumulations can lead to hole or electron blocking effects
which might benefit device performance. In 2009 the difference in vertical
distribution on P3HT:PCBM solar cells was shown to cause problems with
electron mobility which ends up with the yielding of very poor device
efficiencies.[56] Simple changes to device architecture – spin coating a thin
layer of PCBM on top of the P3HT – greatly enhance cell reproducibility, by
providing reproducible vertical separation between device components. Since
higher contact between the PCBM and the cathode is required for better
efficiencies, this largely increases device reproducibility.

According to neutron scattering analysis, P3HT:PCBM blends have been


described as "rivers" (P3HT regions) interrupted by "streams" (PCBM regions).

15
CHAPTER 5
SOLVENT EFFECTS

Conditions for spin coating and evaporation affect device efficiency. Solvent
and additives influence donor-acceptor morphology. Additives slow down
evaporation, leading to more crystalline polymers and thus improved hole
conductivities and efficiencies. Typical additives include 1,8-octanedithiol,
ortho-dichlorobenzene, 1,8-diiodooctane (DIO), and nitrobenzene. The DIO
effect was attributed to the selective solubilization of PCBM components,
modifies fundamentally the average hopping distance of electrons, and thus
improves electron mobility. Additives can also lead to big increases in
efficiency for polymers. For HXS-1/PCBM solar cells, the effect was correlated
with charge generation, transport and shelf-stability. Other polymers such as
PTTBO also benefit significantly from DIO, achieving PCE values of more than
5% from around 3.7% without the additive.

Polymer Solar Cells fabricated from chloronaphthalene (CN) as a co-solvent


enjoy a higher efficiency than those fabricated from the more conventional pure
chlorobenzene solution. This is because the donor-acceptor morphology
changes, which reduces the phase separation between donor polymer and
fullerene. As a result, this translates into high hole mobilities. Without co-
solvents, large domains of fullerene form, decreasing photovoltaic performance
of the cell due to polymer aggregation in solution. This morphology originates
from the liquid-liquid phase separation during drying; solve evaporation causes
the mixture to enter into the spinodal region, in which there are significant
thermal fluctuations. Large domains prevent electrons from being collected
efficiently (decreasing PCE).

16
Small differences in polymer structure can also lead to significant changes
in crystal packing that inevitably affect device morphology. PCPDTBT differs
from PSBTBT caused by the difference in bridging atom between the two
polymers (C vs. Si), which implies that better morphologies are achievable with
PCPDTBT:PCBM solar cells containing additives as opposed to the Si system
which achieves good morphologies without help from additional substances.

17
CHAPTER 6
CURRENT VOLTAGE BEHAVIOUR

Organic photovoltaics, similar to inorganic photovoltaics, are generally


characterized through current-voltage analysis. This analysis provides multiple
device metrics values that are used to understand device performance. One of
the most crucial metrics is the Power Conversion Efficiency (PCE).

Figure 6.1
PCE (η) is proportional to the product of the short-circuit current (JSC), the
open-circuit voltage (VOC), and the fill factor (FF), all of which can be
determined from a current-voltage curve.

18
Where P(in) is the incident solar power.
The short circuit current (Jsc), is the maximum photocurrent generation
value. It corresponds to the y-intercept value of standard current-voltage curve
in which current is plotted along the y-axis and voltage is plotted along the x-
axis. Within organic solar cells, the short circuit current can be impacted by a
variety of material factors. These include the mobility of charge carriers, the
optical absorption profile and general energetic driving forces that lead to a
more efficient extraction of charge carriers

The open-circuit voltage (Voc) is the voltage when there is no current


running through the device. This corresponds to the x-intercept on a current-
voltage curve. Within bulk heterojunction organic photovoltaic devices, this
value is highly dependent on HOMO and LUMO energy levels and work
functions for the active layer materials

Since power is the product of voltage and current, the maximum power point
occurs when the product between voltage and current is maximized.

The fill factor, FF, can be thought of as the "squareness" of a current voltage
curve. It is the quotient of the maximum power value and the product of the
open-circuit voltage and short circuit current. This is shown in the image above
as the ratio of the area of the yellow rectangle to the greater blue rectangle. For
organic photovoltaics, this fill factor is essentially a measure of how efficiently
generated charges are extracted from the device. This can be thought of as a
"competition" between charges transporting through the device, and charges
that recombine.

19
A major issue surrounding polymer solar cells is the low Power Conversion
Efficiency (PCE) of fabricated cells. In order to be considered commercially
viable, PSCs must be able to achieve at least 10–15% efficiency this is already
much lower than inorganic PVs. However, due to the low cost of polymer solar
cells, a 10–15% efficiency is commercially viable.

Recent advances in polymer solar cell performance have resulted from


compressing the bandgap to enhance short-circuit current while lowering the
Highest Occupied Molecular Orbital (HOMO) to increase open-circuit voltage.
However, PSCs still suffer from low fill factors (typically below 70%).
However, as of 2013, researchers have been able to fabricate PSCs with fill
factors of over 75%. Scientists have been able to accomplish via an inverted
BHJ and by using nonconventional donor / acceptor combinations.

20
CHAPTER 7
COMMERCIALIZATION

Polymer solar cells have yet to commercially compete with silicon solar
cells and other thin-film cells. The present efficiency of polymer solar cells lies
near 10%, well below silicon cells. Polymer solar cells also suffer from
environmental degradation, lacking effective protective coatings.

Further improvements in performance are needed to promote charge carrier


diffusion; transport must be enhanced through control of order and morphology;
and interface engineering must be applied to the problem of charge transfer
across interfaces.

Figure 7.1

Research is being conducted into using tandem architecture in order to


increase efficiency of polymer solar cells. Similar to inorganic tandem
architecture, organic tandem architecture is expected to increase efficiency.
21
Compared with a single-junction device using low-bandgap materials, the
tandem structure can reduce heat loss during photon-to-electron conversion.

Polymer solar cells are not widely produced commercially. Starting in


2008‘Konarka Technologies started production of polymer-fullerene solar cells.
The initial modules were 3–5% efficient, and only last for a few years. Konarka
has since filed for bankruptcy, as those polymer solar cells were unable to
penetrate the PV market.

PSCs also still suffer from low fill factors (typically below 70%). However,
as of 2013, researchers have been able to fabricate PSCs with fill factors of over
75%. Scientists have been able to accomplish via an inverted BHJ and by using
nonconventional donor / acceptor combinations.

However, efforts are being made to upscale manufacturing of polymer solar


cells, in order to decrease costs and also advocate for a practical approach for
PSC production. Such efforts include full roll-to-roll solution processing.
However, roll-to-roll solution processing is ill-suited for on-grid electricity
production due to the short lifetime of polymer solar cells. Therefore,
commercial applications for polymer solar cells still include primarily consumer
electronics and home appliances.

CHAPTER 8
22
CHALLENGES
Understanding the mechanical properties of organic semiconductors and in
particular, the wide range of failure mechanisms of operating organic solar cell
devices are critical in determining the operational stability of organic solar cells
for various applications.The mechanical properties of organic solar cells can be
attributed to intermolecular and surface forces present in the material. These
attributes are not only influenced by the molecular structure but are also quite
sensitive to processing conditions, making the study of mechanical properties of
polymer thin films such as tensile modulus, ductility and fracture toughness
under strain rather difficult. Because of this, it is nontrivial to quantify a "figure
of merit" that will predict the mechanical stability of a device and the robustness
of a device under strain will depend on many factors.

Most often, the substrate provides support to the device and mechanical
failure of the substrate will lead to suboptimal power conversion efficiency of
the device. Hence while it is necessary that the substrate provides mechanical
support to the organic active layer, care must be taken to ensure that increasing
the tensile strength of the substrate does not come at the cost of the film
fracturing at low strains. In general, it is desirable that the active layer deforms
in tandem with the substrate. This is made possible with a low elastic modulus
and high elastic limit. The ductility of a thin film is commonly measured as the
strain at which cracks appear on the film. However, the crack onset strain is also
dependent on other factors such as the degree of cohesion/adhesion between the
film and the substrate. Various studies have related the cohesive or adhesive
fracture energy Gc , defined as the work required to break separate polymer
interfaces to molecular parameters and processing conditions. Along with the

23
cohesion, the trajectory of crack propagation following formation depends on
the mechanical properties of the material the crack propagates through. In
polymers like P3HT that exhibit good plasticity, a plastic zone forms at the
crack tip upon the application of a tensile strain normal to the plane of the
device and expands until it is confined by either crystalline domains in the film
or by a rigid substrate, thus dissipating the deformation energy and decreasing
the cohesion between the interfaces. The mechanical buckling technique has
also proven to be quite successful in determining the elastic moduli of various
organic thin films. The method is based on the buckling instability that gives
rise to wrinkles in the film under a compressive strain. The wavelength of the
wrinkling pattern, can be related to the tensile modulus of the film in terms of
the film thickness and elastic modulus of the substrate.

In the design of devices incorporating organic solar cells, Gc and strain at


fracture have been identified as two metrics that are important to consider. The
bulk heterojunction (BHJ) layer is usually the weakest layer of an organic solar
cell, so it is necessary to design the BHJ materials to be mechanically stable,
with a target Gc of 5 J m-2 and a target strain at fracture of 20-30%. Polymer-
based acceptors have been shown to exhibit superior mechanical properties
when compared to small-molecule acceptors and fullerene-based acceptors.
Additionally, the mechanical properties of polymer-based acceptors are
influenced by Mn, the number-average molecular weight of the polymer
molecules. It was determined that mechanical properties increase with
increasing Mn, but only once Mn has surpassed Mc, the critical molecular
weight at which entanglements cause the rate of viscosity change to increase
with increasing Mn. This phenomenon occurs because the rate of chain
entanglement and miscibility between the polymer acceptor and donor both

24
increase. The effect of these characteristics is that the plastic deformation of
these materials in reaction to mechanical stress was high, meaning that more of
the energy was dissipated, while the materials with less mechanical strength
fractured more readily.

CHAPTER 9
25
CONCLUSION

One major area of current research is the use of non-fullerene acceptors.


While fullerene acceptors have been the standard for most organic photovoltaics
due to their compatibility within bulk heterojunction cell designs as well as their
good transport properties, they do have some fallbacks that are leading
researchers to attempt to find alternatives. Some negatives of fullerene
acceptors include their instability, that they are somewhat limited in energy-
tunability and they have poor optical absorption. Researchers have developed
small molecule acceptors that due to their good energy tunability, can exhibit
high open-circuit voltages. Combining a polymer donor (D18) with a small
molecule acceptor (Y6), scientists have fabricated organic solar cells in the
laboratory giving high efficiencies over 18%. However, there are still major
challenges with non-fullerene acceptors, including the low charge carrier
mobilities of small molecule acceptors, and that the sheer number of possible
molecules is overwhelming for the research community.

A challenge facing the development of organic solar cells utilizing non-


fullerene acceptors (NFAs) is the selection of a solvent that has a high boiling
point and is environmentally friendly, whereas conventional solvents such as
chloroform (CF) tend to exhibit low boiling points and toxicity. Such a solvent
is required for further scale-up of organic solar cells, but has also been
associated with decreases in PCE due to poor solubility of donor and acceptor
materials within the solvent. Appending alkyl chains to NFAs has led to
increases in solubility but decreases in molecular packing (π-stacking), which
leads to no net impact on PCE. The use of guest assistance has been found to
benefit both solubility and molecular packing. A guest molecule named BTO

26
with oligo(ethylene glycol) (OEG) side chains used in conjunction with the
NFA Y6 as the acceptor, PM6 as the donor, and paraxylene (PX) as the high-
melting-point and sustainable solvent led to an increase in PCE from 11% to
over 16%, regarded an acceptable level of efficiency. A further modification
that has been successful in the development of cleaner organic photovoltaics is
the hot-spin coating of substrates by non-halogenated solvents. It was found that
the temperature at which hot-spin coating was operated altered the solution to
solid phase evolution of the acceptor-donor blends so that higher temperatures
resulted in a higher acceptor concentration in the surface of the substrate. This is
because higher temperatures facilitated decreased aggregation and precipitation,
allowing the substrate to retain a higher acceptor concentration. In an
experiment, organic solar cells constructed with ternary blends of PM6 donor
and Y6-1O and BO-4Cl acceptors and various non-halogenated solvents
including o-xylene and toluene exhibited PCE values of over 18%, which are
the most efficient organic photovoltaics constructed with non-halogenated
solvents, to date. Further morphological analyses showed that the hot-spun
OPVs prepared with non-halogenated solvents exhibited similar morphological
characteristics to that of OPVs prepared with halogenated solvents.

Small molecules are also being heavily researched to act as donor materials,
potentially replacing polymeric donors. Since small molecules do not vary in
molecular weights the way polymers do, they would require less purification
steps and are less susceptible to macromolecule defects and kinks that can
create trap states leading to recombination. Recent research has shown that
high-performing small molecular donor structures tend to have planar 2-D
structures and can aggregate or self assemble. Sine performance of these
devices is highly depended on active layer morphology, present research is

27
continuing to investigate small molecule possibilities, and optimize device
morphology through processes such as annealing for various materials.

REFERENCES
1) Ameri, Tayebeh; Dennler, Gilles; Lungenschmied, Christoph; Brabec,
Christoph (2009). "Organic tandem solar cells: A review". Energy &

28
Environmental Science. 2 (4): 348. doi:10.1039/B817952B. Retrieved
2019-05-20.

2) Pulfrey, L.D. (1978). Photovoltaic Power Generation. New York: Van


Nostrand Reinhold Co. ISBN 9780442266400.

3) Nelson, Jenny (2011-10-01). "Polymer:fullerene bulk heterojunction solar


cells". Materials Today. 14 (10): 462–470. doi:10.1016/S1369-
7021(11)70210-3.

4) "What can organic solar cells bring to the table?". Retrieved 26 March
2021.

5) Luther, Joachim; Nast, Michael; Fisch, M. Norbert; Christoffers, Dirk;


Pfisterer, Fritz; Meissner, Dieter; Nitsch, Joachim (2000). "Solar
Technology". Ullmann's Encyclopedia of Industrial Chemistry.
doi:10.1002/14356007.a24_369. ISBN 3527306730.

6) Jørgensen, Mikkel; Norrman, Kion; Krebs, Frederik C. (2008).


"Stability/degradation of polymer solar cells". Solar Energy Materials and
Solar Cells. 92 (7): 686. doi:10.1016/j.solmat.2008.01.005.

7) Po, Riccardo; Carbonera, Chiara; Bernardi, Andrea; Tinti, Francesca;


Camaioni, Nadia (2012). "Polymer- and carbon-based electrodes for
polymer solar cells: Toward low-cost, continuous fabrication over large
area". Solar Energy Materials and Solar Cells. 100: 97.
doi:10.1016/j.solmat.2011.12.022.

8) Scharber, M. C.; Mühlbacher, D.; Koppe, M.; Denk, P.; Waldauf, C.;
Heeger, A. J.; Brabec, C. J. (2006). "Design Rules for Donors in Bulk-
Heterojunction Solar Cells—Towards 10 % Energy-Conversion
Efficiency" (PDF). Advanced Materials. 18 (6): 789.
Bibcode:2006AdM....18..789S. doi:10.1002/adma.200501717. S2CID
13842344.

9) You, Jingbi; Dou, Letian; Yoshimura, Ken; Kato, Takehito; Ohya,


Kenichiro; Moriarty, Tom; Emery, Keith; Chen, Chun-Chao (5 February
2013). "A polymer tandem solar cell with 10.6% power conversion
efficiency". Nature Communications. 4: 1446.
Bibcode:2013NatCo...4.1446Y. doi:10.1038/ncomms2411. PMC
3660643. PMID 23385590.

29
30

You might also like